THE ISOPERIMETRIC INEQUALITY ON COMPACT RANK ONE SYMMETRIC SPACES AND BEYOND

YASHAR MEMARIAN

Abstract. Klartag’s needle decomposition technique enables one to obtain strong isoperimetric inequalities on Riemannian manifolds other than the clas- sical known examples. As a result, in this paper, we obtain sharp isoperimetric inequalities for compact rank one symmetric spaces (CROSS). Namely, for the n real projective space RP , we demonstrate that the isoperimetric regions are k n given by either the geodesic balls or tubes around some RP ⊂ RP . For the n CP , the isoperimetric regions are given by either k n the geodesic balls or tubes around some CP ⊂ CP . And for the quaternionic projective space, the isoperimetric regions are given by either the geodesic balls k n or tubes around some HP ⊂ HP .

1. Introduction Isoperimetric problems are some of the oldest problems in geometry. Given a space, one looks for domains of a given volume with the least boundary area. It is known that in model spaces (i.e. Euclidean spaces, spheres and hyperbolic spaces), for every given number v, the intrinsic balls with volume v have the least surface area among every domain with the same volume. When we leave the world of model spaces, or when we are dealing with geometric spaces with boundary, the solution to isoperimetric problems has provided real difficulties. Of course, when the number v is small enough, for spaces which locally look like Euclidean spaces (for instance manifolds), one expects that the solution of isoperimetric problems still would be the metric balls. However, for larger v and shapes with non-constant curvature or with non-smooth boundaries, the isoperimetric problem in general is very hard to solve. There are several books and surveys related to isoperimetric problems, for instance [23], [21], [15], [6], [3]. In [11], Gromov-Milman prove a general isoperimetric problem on spheres with non-necessarily canonical Riemannian volume. They use their result to obtain an isoperimetric inequality for unit spheres of uniformly-convex Banach spaces (for arXiv:1710.03952v3 [math.MG] 11 Jul 2021 example Lp-unit spheres). Their method relies on the geometry and topology of the sphere. They use a powerful technique (known as the localization method) in order to prove their result(s). Obtaining similar results for more general manifolds using the localization method was not possible since one did not have such tools

Date: June 2020. 2010 Mathematics Subject Classification. 53C21. Key words and phrases. Isoperimetry, needle, Ricci Curvature, , Positively Curved Manifolds . We thank Karsten Grove for motivating us to write this paper. The paper started when the author visited the University of Notre-Dame. We would like to thank the hospitality of the mathematics department during our stay. We also thank the referees for their extremely helpful comments which made this paper realizable. 1 2 YASHAR MEMARIAN for spaces other than model spaces (even on the canonical hyperbolic space we are not aware of any result for which the localization method has been used). Recently, Klartag in [18] proves a localization theorem on every closed Riemann- ian manifold. The topic of this paper concerns utilizing Klartag’s localization results in order to prove sharp isoperimetric inequalities for compact rank one symmetric spaces, namely the real, complex and quaternionic projective spaces. We begin by recalling Klartag’s results on needle decomposition and its link to the isoperimetric problems. In further sections we concentrate on studying the isoperimetric problem on real, complex and quaternionic projective spaces.

2. Klartag’s Generalisation of Gromov-Milman Isoperimetric Inequality on Riemannian Manifolds In this manuscript, we are mainly interested in closed, smooth n-dimensional Riemannian manifolds with some lower bounds on the Ricci curvature. A Rie- mannian manifold M satisfies the curvature-dimension condition CD(k, N), if the dimension of M is at most equal to N and if for every point m ∈ M and every tangent vector v ∈ TmM we have: (2.1) Ricci(v, v) ≥ k.g(v, v), where g is the metric tensor of M and Ricci is the Ricci tensor of M. When for a Riemannian manifold M, there exists a k ∈ R such that for every point m ∈ M and every tangent vector v ∈ TmM, inequality (2.1) holds, we simply write that on M we have: Ricci ≥ k. The curvature-dimension condition and its extension to the case of weighted Rie- mannian manifolds were first defined in the pioneer work [1]. The volume Riemannian volume of M is denoted by voln emphasizing the di- mension of M. We recall the (metric-measure) invariant separation distance introduced by Gro- mov in [13] (and motivated by ideas in [11]).

Definition 2.1 (Separation Distance). Let κ1, κ2 > 0 be such that

κ1 + κ2 < 1.

The separation distance on M (with respect to κ1 and κ2), denoted by:

Sep(M, κ1, κ2), is the supremum of those δ, where for i = 1, 2, subsets Ai ⊂ M exist with µ(Ai) ≥ κi and dist(A1,A2) ≥ δ. Here

dist(A1,A2) = inf d(a1, a2), (a1,a2)∈A1×A2 where d(.) is the distance induced by the Riemannian metric on M. We say that the sets Ai (for i = 1, 2) with µ(Ai) = κi, realize the separation distance if dist(A1,A2) = Sep(M, κ1, κ2).

Remark 2.1. • We denote the distance between two sets A1,A2 ⊂ M defined in definition 2.1 by:

d(A1,A2) = dist(A1,A2). This distance should not be confused with the Hausdorff distance. THE ISOPERIMETRIC INEQUALITY ON COMPACT RANK ONE SYMMETRIC SPACES 3

• One can define the separation distance with respect to several (positive real) numbers κ1, ··· , κk, in the same manner as defined with respect to two numbers. For the purpose of this paper, we only require this definition with respect to two numbers. We now need the definition of a needle on n-dimensional Riemannian manifolds M with Ricci(M) ≥ Ricci(Sn). Since we are using Klartag’s work on needle decom- position of Riemannian manifolds, we present the definition of the needle exactly as it is defined in [18]: Definition 2.2 (Needles). Let M be an n-dimensional Riemannian manifold. A CD(k, n)-needle on M is a measure ν on M such that the following holds: • There exists a non-empty open set A ⊂ R and a smooth function ψ : A → R. • The function ψ1/n−1 satisfies the following differential inequality on A:

1 00 k 1 (2.2) (ψ n−1 ) + ( )ψ n−1 ≤ 0. n − 1 • There exists a minimizing geodesic γ : A → M such that:

ν = γ∗(ψ(x)dx). Hence the measure ν defined on M is the push-forward of the measure ψ(x)dx on A which has a density function ψ with respect to the Lebesgue measure dx. A smooth function satisfying inequality (resp. equality) (2.2) for k = n−1 is called a sinn−1-concave (resp. affine) function. A measure ν which has a sinn−1 density function is called a sinn−1-concave measure. Remark 2.2. • Sometimes it may be important to recall the parameter in- terval A on which the pull-back of a needle is defined and supported. In such case, we may denote a CD(k, n)-needle by (I, ν) where I ⊂ R is a non-empty open connected subset of R and the pull-back of the measure ν is supported on I. This is useful when one considers needles on intervals of R i.e. geodesic segments of R. • In definition 2.2, when k = n − 1, the needles may be referred to as sinn−1- concave needles (or measures). • The inequality (2.2) is a local definition of the sinn-concave functions. One n (equivalently) may define a sin -concave function on an interval A ⊂ R as a function satisfying the following inequality: x + x f(x )1/n + f(x )1/n (2.3) f( 1 2 )1/n ≥ 1 2 , 2 |x2−x1| 2 cos( 2 )

for every x1, x2 ∈ A. See [26] for a proof of the equivalence between in- equalities (2.3) and (2.2). • According to definition 2.2, we may view needles as intrinsic metric-measure spaces (i.e. not necessarily embedded on a manifold M) where the geomet- ric space is an interval of R, the metric being the Euclidean metric and the measure being the sinn-concave probability measure defined on this inter- val. On the other hand, it is important to bear in mind that needles inherit the geometry of the manifold on which they are embedded. 4 YASHAR MEMARIAN

The following definition is handy since we shall be dealing with probability mea- sures all along this manuscript: Definition 2.3 (Normalizing Constant). Let f(x)dx be a measure supported on an interval A ⊂ R where dx is the Lebesgue measure. We say C ∈ R is the normalizing constant for f(x)dx, if Cf(x)dx is a probability measure supported on A. We present some classical examples to get familiarized with some needles:

Example 2.1. On the canonical sphere Sn, let {x, −x} be two diametrically op- posite points. Let γ be a maximal geodesic from x to −x. This geodesic being parametrized by the interval (0, π) which is enhanced with the normalized measure C cos(t)n−1dt. Then the push-forward of this measure on Sn is an example of a needle on the canonical sphere. Example 2.2. A sinn-affine probability measure ν defined on an open connected n interval A ⊂ R is defined by ν = (C1 sin(t) + C2 cos(t)) dt for some C1,C2 ∈ R where dt is the Lebesgue measure. We have mentioned that one could consider needles defined intrinsically, and viewing them as such a needle is a metric-measure space defined on its own. But how are the needles constructed on a manifold M (with Ricci(M) ≥ Ricci(Sn)? For the case M = Sn, the construction of a needle is explained in [11]. Indeed, the authors show that needles are obtained as a limit of S1 ⊃ S2 ⊃ · · · , where each n Si is a (geodesically) convex subset of S . Here, the limit is a weak limit of the normalized volume of the Si (for i = 1 · · · ∞). When M is no longer the canonical sphere, the definition and construction of needles are explained in [18]. The strategy of the author in [18] is to define some geometric objects called needle candidates, and then show that every needle can- didate is a needle in the sense of definition 2.2. We present the definition of the needle candidates below and shall understand needles as some measures which are tied to the geometry of M. Let γ : [0, a] → M be a minimizing geodesic on M which is parametrized by its arclength. One may use the Riemannian connection on M and define the covariant derivative of any smooth vector field J along γ. A smooth vector field J along γ is a smooth vector field J such that J(t) ∈ Tγ(t)M (the tangent space of M at point γ(t)). The covariant derivative of J along γ is denoted by 0 J = ∇γ0 J, where γ0 is the tangent vector field along γ. A vector field along γ is called a Jacobi field if the following equation is satisfied for every t ∈ [0, a]: J 00(t) = R(γ0(t),J(t))γ0(t), where R is the Riemann curvature tensor and for X,Y,Z ∈ TM is defined by:

R(X,Y )Z = ∇X ∇Y Z − ∇Y ∇X Z − ∇[X,Y ]Z. For every smooth function f ∈ C∞(M), [X,Y ] is defined by [X,Y ](f) = X(Y (f)) − Y (X(f)). With this background, we are now ready to give the definition of needle candi- dates presented in [18]: THE ISOPERIMETRIC INEQUALITY ON COMPACT RANK ONE SYMMETRIC SPACES 5

Definition 2.4 (Needle Candidates). A measure ν on a Riemannian manifold M is called a needle candidate on M if there exists a non-empty subset (a, b) ⊂ R with a, b ∈ R ∪ {−∞, +∞}, a measure µ on (a, b), a minimizing geodesic γ :(a, b) → M and Jacobi fields J1(t), ··· ,Jn−1(t) along γ with the following properties: • The measure ν is the push-forward of the measure µ under the map γ. 0 • Denote Jn = γ . Then the measure µ is absolutely continuous with respect to the Lebesgue measure in (a, b) and its density is proportional to q t → det(< Ji(t),Jk(t) >)i,k=1,··· ,n. • There exists t ∈ (a, b) with 0 0 0 < Ji(t), γ (t) > = < Ji (t), γ (t) >= 0 (i = 1 ··· , n − 1), and 0 0 < Ji (t),Jk(t) > = < Jk(t),Ji(t) > (i, k = 1, ··· , n − 1). • Either for all t ∈ (a, b) the vectors

J1(t), ··· Jn−1(t) ∈ Tγ(t)M are linearly independent or else for all t ∈ (a, b) these vectors are linearly dependent. Remark 2.3. In [18] where a needle candidate is defined, there exists a Lipschitz function u such that the needle candidate is defined along the integral curves of the gradient of u. We refer the reader to [18] for all details. The following, proved in [18], relates the needle candidates to the needles of definition 2.2:

Proposition 2.1. Let M be a Riemannian manifold with Ricci(M) ≥ Ricci(Sn). Let ν be a needle candidate defined in 2.4. Then ν is a needle as defined in 2.2. We would like to extend the definition of separation distance to the class of needles and/or needle candidates. Should one follow the intrinsic point of view, one sets n, k ≥ 0 and considers the class of CD(k, n)-needles. We consider probability measures and then this is a cone in the space of probability measures with support in R (see [12]). Let A be a class of CD(k, n)-(probability) needle. Then one defines the needle separation distance with respect to A as follows:

Definition 2.5. Given two positive real numbers κ1, κ2 > 0, such that

κ1 + κ2 < 1, the needle separation distance with respect to A is defined by

NA(κ1, κ2) = sup Sep(ν, κ1, κ2), ν∈A

where Sep(ν, κ1, κ2) is the separation distance on the metric-measure space (I, ν), and the supremum is taken over every needle in A. We say ν realizes the needle separation distance NA(κ1, κ2) if

Sep(ν, κ1, κ2) = NA(κ1, κ2). 6 YASHAR MEMARIAN

Moreover the sets Ii ⊆ I with ν(Ii) = κi (for i = 1, 2) realize the needle separation distance if

d(I1,I2) = NA(κ1, κ2). We are interested in studying the isoperimetric problem on a Riemannian man- ifold. For this purpose we need the separation distance for the needle candidates, as the latter reflects in some ways the isoperimetric properties of M. Definition 2.6 (Needle Candidate Separation Distance). Let M be a Riemannian manifold of dimension n with Ricci(M) ≥ Ricci(Sn). Given two positive real numbers κ1, κ2 > 0 such that

κ1 + κ2 < 1, the needle candidate separation distance is defined as follows:

NM (κ1, κ2) = sup Sep(ν, κ1, κ2). ν

Here, Sep(ν, κ1, κ2) is the separation distance on the metric-measure space (γ, ν) (i.e. a geodesic segment and a probability measure defined upon it). The supremum is taken over all needle candidates ν which can be defined on M. We say Ii ⊆ I with ν(Ii) = κi (for i = 1, 2) realize the needle candidate separation distance if

d(I1,I2) = NM (κ1, κ2).

Furthermore we say Ai ⊆ M with µ(Ai) = κi realize the needle candidate separation distance if

d(A1,A2) = NM (κ1, κ2). In the figure below, we review the definition of the separation and needle sepa- ration distances on an example:

In the upper part, one can find the graph of the function C cos(t)n on the in- terval (−π/2, +π/2), for some n > 1 where C is the normalizing constant. On the left and right hand sides of the interval (−π/2, +π/2), we find two sub-intervals with two given measures (say κ1 and κ2). The separation distance for the needle ((−π/2, +π/2),C cos(t)ndt) is the distance between these two intervals where the distance is defined in definition 2.1 (and should not be confused with the Hausdorff THE ISOPERIMETRIC INEQUALITY ON COMPACT RANK ONE SYMMETRIC SPACES 7 distance). Below the graph of this needle, one can observe a disc which is seen as the projection of a (hemi)-sphere. The diameter of the sphere is equal to the diameter of the needle pictured above it. The left and right hand sides of this disc are the projection of two spherical balls. Each spherical ball has a measure equal to the measure of the interval on the needle shown above it. The separation distance for this sphere is now defined to be the distance (in the sense of definition 2.1) between these two spherical balls. We are now ready to present the following strong Theorem which is due to Klartag: Theorem 2.1. Let M be a closed, smooth Riemannian manifold where Ricci(M) ≥ n Ricci(S ). Given κ1, κ2 > 0 such that

κ1 + κ2 < 1, we have:

Sep(M, κ1, κ2) ≤ NM (κ1, κ2). Similar to the proof in [11], this Theorem is proved by using the localization technique. Gromov-Milman prove this theorem on the sphere Sn, in which the topology of the sphere enables one to construct such needles. However, on general Riemannian manifolds, one can not provide needle decomposition in the same way as one pro- vides needle decomposition on Sn. To remedy this issue, Klartag (in [18]) presents a localization theorem on every closed Riemannian manifold. Theorem 2.1 has very important consequences, when applied on specific examples. We provide a proof of Klartag’s separation distance Theorem 2.1 here and further present a corollary of this Theorem which links this result to the study of (sharp) isoperimetric inequalities on Riemannian manifolds. Theorem 2.2 (Klartag). Let M be a closed, smooth Riemannian manifold with Ricci(M) ≥ Ricci(Sn). Let µ be the Riemannian volume on M. Let f be a µ- integrable function such that Z f(x)dµ(x) = 0. M

Then, there exist a partition Π of M, a measure ν on Π, a family {µπ}π∈Π of measures on M such that: • For any Lebesgue-measurable set U ⊆ M we have: Z µ(A) = µπdν(π). Π • For ν-almost every π ∈ Π, the set π ⊂ M is the image of a minimizing geodesic. The measure µπ is supported on π, and either π is a singleton or a needle candidate in the sense of definition 2.4. • For ν-almost any π ∈ Π we have: Z fdµπ = 0. π Remark 2.4. Consult [18] for a precise understanding on how the partitions are defined and constructed. In [18] Klartag proves that the partition of Theorem 2.2 is obtained as the integral curves of the gradient of some function u : M → R. 8 YASHAR MEMARIAN

2.1. Proof of Theorem 2.1.

Proof. Let κ1, κ2 > 0 such that

κ1 + κ2 < 1 be given. Since M is compact, there exist Ui which realize the separation distance (with respect to κi) for i = 1, 2. According to Theorem 2.2, there exists a partition of M into needle candidates such that for almost every needle candidate ν we have:

ν(Ui) = κi.

Indeed, assume µ(U1) = κµ(U2). Hence the integral of the function ξU1 − κξU2 is equal to zero on M. Here ξ is the indicatrice function. We apply Theorem 2.2 with respect to this function. Since we have a partition of M into needle candidates (which satisfy the measure assumption above), there exists (at least) a needle ν in this partition such that

d((I ∩ U1), (I ∩ U2)) ≥ d(U1,U2), where γ : A → M and γ(A) = I ⊂ M. Hence by definition we have:

NM (κ1, κ2) ≥ d((I ∩ U1), (I ∩ U2))

≥ d(U1,U2)

= Sep(M, κ1, κ2). This completes the proof of Theorem 2.1.  Remark 2.5. Theorem 2.1 is valid for the class of weighted Riemannian manifolds satisfying the generalized curvature-dimension property. Consult [18] for this pur- pose. Notation 1. For A ⊂ M and for ε > 0, A + ε stands for the ε-neighborhood of the set A. Let us see how Theorem 2.1 may be used for solving isoperimetric (type) prob- lems. This is explained in [18] and [19]. Proposition 2.2. Let M be a smooth closed n-dimensional Riemannian manifold with Ricci(M) ≥ Ricci(Sn). Let µ be the normalized Riemannian volume and ε > 0 be fixed. For every 0 < v ≤ 1, and for every open set A with measure equal to v, let w(A) = 1 − µ(A + ε).

Suppose (B,B1) ⊆ M × M exists which realizes the needle separation distance N(v, w(A)). Then we have: µ(A + ε) ≥ µ(B + ε).

Proof. Assume A1 is an open set on M with a measure equal to v. Let ε > 0 be given and let A2 = M \ (A1 + ε) and A3 = (A1 + ε) \ A1. Let µ(A2) = w. According to Theorem 2.1 we have:

Sep(M, v, w) ≤ NM (v, w).

By assumption, let (B1,B2) ⊆ M × M be such that µ(B1) = v, µ(B2) = w and

d(B1,B2) = NM (v, w). THE ISOPERIMETRIC INEQUALITY ON COMPACT RANK ONE SYMMETRIC SPACES 9

Therefore we obtain:

d(A1,A2) = ε

≤ d(B1,B2).

This means µ(M \ (B1 + ε)) ≥ µ(A2). And thus

µ(A1 + ε) ≥ µ(B1 + ε). The proof therefore follows.  3. Isoperimetric properties of Needles

Considering the CD(k, n)-needles as metric-measure spaces of their own, there is the natural problem of characterizing the needle(s) which attains N(κ1, κ2) for a given κ1, κ2 ∈ (0, 1). This question only makes sense if we define A (the domain in the cone of CD(k, n)-needles) on which we take the supremum of Sep(ν, κ1, κ2). Here we will be dealing with the most general isoperimetric properties of needles. In other words, given n ∈ N, we will be studying N(κ1, κ2) when A is the entire cone of CD(k, n)-needles. Most of the techniques which follow shall also be used to study NM (κ1, κ2) for some specific (symmetric) Riemannian manifolds. Studying the properties of a needle is equivalent to studying the properties of their densities of their pull-back on an interval of R. For simplicity, we take k = n−1 here and present some properties of the sinn-concave functions (measures). There exists a classical way to study the properties of a sinn-concave function which is related to the local definition of such functions, namely the differential inequality (2.2). This is provided by the Sturm-Liouville theory (about differential equations): Let u satisfy inequality (2.2) (resp. inequality) on the interval [a, b]. Let v satisfy the equality in (2.2) (resp. equality) on the same interval [a, b], where u(a) = v(a) and u(b) = v(b) (or u(a) = v(a) and u0(a) = v0(a)). Then u ≤ v on [a, b]. In order to obtain integral formulae using the above method, one may use the following Lemma (assigned to Gromov and used to prove the Bishop-Gromov in- equality):

Lemma 3.1. Let f, g be real positive function on [0,A] where A ∈ R+ such that f g is a non-increasing function, then Z x f(t)dt 0 Z x g(t)dt 0 is a non-increasing function on [0,A]. There exists a more geometric way (related to the global properties of sinn- concave functions, namely inequality (2.3)) to study the properties of sinn-concave functions. This is worth explaining here: • Let µ be a measure which has a density function f and suppose f is a sinn-concave function. By definition the function g = f 1/n is a sin-concave function. 10 YASHAR MEMARIAN

• Consider the support of µ and suppose the support is I ⊆ [0, π]. 1 • Transport I on the (half)-unit circle (denoted by S+), which is the unit circle in R2 with non-negative y-coordinates. By transport we mean a unit 1 speed parametrization map I → S+. 2 • Consider the cone over the arc I in R+ denoted by co(I), defined by: co(I) = {∪(tI) | 0 ≤ t ≤ 1}. • The definition of a sin-concave function can now be given as follows: there exists a function G in C(I) such that G is a concave function and G is 1-homogeneous (i.e. G(λx) = λx for λ ≥ 0) and G(x) = g(x) for x ∈ I. • Therefore, to understand the properties of a sin-concave funcion/measure, one could study the properties of its extension on a cone in R2 where the extended function is a 1-homogeneous concave function. The above approach is followed in [20], and most of the properties of sinn-concave measures are obtained in this way. We present the main properties in the following Lemma and omit the proof. For complete proof one could consult [20]. Lemma 3.2. • Let 0 < ε ≤ π/2. Let τ ≥ ε. Let f be a non-negative sinn- concave function on [0, τ], which attains its maximum at 0. Let h(t) = C cos(t)n where C is chosen such that f(ε) = h(ε). – Then we have ( f(x) ≥ h(x) for x ∈ [0, ε], f(x) ≤ h(x) for x ∈ [ε, τ]. And τ ≤ π/2. – For every k ≥ 0 and ε ≤ π/2, we have : Z min{ε,τ} Z ε f(t) sin(t)kdt cos(t)n sin(t)kdt (3.1) 0 ≥ 0 . Z τ π/2 k Z f(t) sin(t) dt cos(t)n sin(t)kdt 0 0 k • Let (s, k) ∈ N × N be such that s ≤ k. Then every sin -concave function is also a sins-concave function. Moreover, if f is sinm-concave, g is sinn- concave, then fg is sinm+n-concave. • The above properties hold for functions which satisfy the differential inequal- ity of (2.2). Indeed, for such functions, everything above is to compare with model functions: √ C sinn( λt) √ , λ which are the functions satisfying the differential equality in (2.2). • Let I ⊂ R be a connected closed interval. Let µ = f(t)dt be a probability n measure where f is a sin -concave function. Let κ1, κ2 > 0 be such that

κ1 + κ2 < 1. Then, the separation distance

Sep(µ, κ1, κ2), THE ISOPERIMETRIC INEQUALITY ON COMPACT RANK ONE SYMMETRIC SPACES 11

is realized by A1,A2 ⊂ I with

µ(A1) = κ1

µ(A2) = κ2.

Moreover, both A1 and A2 are connected intervals. n • Let µ1 = C1f(t)dt be a probability measure on (0,L1) where f is a sin - concave function. Let µ2 = C2f(t)dt be a probability measure on (0,L2) where

L2 ≥ L1.

Then for every κ1, κ2 > 0 such that

κ1 + κ2 < 1, we have:

(3.2) Sep(µ2, κ1, κ2) ≥ Sep(µ1, κ1, κ2). Proof. • For the proof of parts 1 to 3, one may consult [20], [12] and [26]. • The existence and connectivity of the intervals which realize the separation distance is obtained from a classical compacity argument followed by the concavity properties of the sinn-concave functions. • For the last part, let 0 < ε ≤ δ. We use Lemma 3.1 to obtain: Z ε Z δ f(t)dt f(t)dt 0 ≥ 0 . Z L1 Z L2 f(t)dt f(t)dt 0 0 By using the above inequality the proof of (3.2) follows.  The general isoperimetric property of CD(n − 1, n)-needles is presented in the following:

Proposition 3.1. Let κ1, κ2 > 0 be such that

κ1 + κ2 < 1. Let µ = f(t)dt be a probability measure where f is a sinn-concave function and µ is supported on the interval I ⊂ R. Then n Sep(µ, κ1, κ2) ≤ Sep(C sin(t) dt, κ1, κ2), where C is the normalizing constant and the probability measure C sin(t)ndt is sup- ported on the interval (0, π). Proof. From Lemma 3.2, we know that the length of the support of the measure µ is at most equal to π. Therefore, we translate (i.e change of variable t → t + a) the function f such that the point on which the (unique) maximum is attained, 12 YASHAR MEMARIAN coincides with the point π/2 on which the maximum of sin(t)n is attained. Let ε > 0 be such that: Z t2 sin(t)ndt t1 Z π = 1 − (κ1 + κ2), sin(t)ndt 0 where t1 = π/2 − ε and t2 = π/2 + ε. From Lemma 3.2 namely equation (3.1) (when one takes k = 0 in this formula), we have: Z t2 sin(t)ndt t1 1 − (κ1 + κ2) = Z π sin(t)ndt 0 Z t2 f(t + a)dt t1 ≤ Z π . f(t + a)dt 0

Hence, there exists a ε1 ≤ ε for which we have Z s2 f(t + a)dt s1 Z π = 1 − (κ1 + κ2), f(t + a)dt 0 where s1 = π/2 − ε1 and s2 = π/2 + ε1. This proves that: n Sep(µ, κ1, κ2) ≤ Sep(C sin(t) dt, κ1, κ2). This ends the proof of the proposition.  3.1. Recollection on Spherical Isoperimetric Problem and the Levy-Gromov Isoperimetric inequality. With the tools presented in the previous sections, namely Proposition 2.2 and Proposition 3.1, we can already recover the sharp isoperimetric inequality on the canonical sphere as well as the Levy-Gromov The- orem. This is explained in [18] and following Klartag’s work, we present these two results here. One could enjoy the powerful localization method by observing how easily one obtains these isoperimetric inequalities. It is interesting to compare the following proofs with the original proofs in which more sophisticated ideas are used. We start with the canonical sphere Sn. The solution of the isoperimetric in- equality on the sphere has been known for many years. Similar to the Euclidean counter-part, there are several different proofs for this result. The proof which uses localization is presented in [11]. We recall this proof here:

Theorem 3.1 (Isoperimetry on the Sphere). Let Sn be the canonical sphere. Let voln denotes the (canonical) Riemannian volume. For every open set A and for every ε > 0 we have:

voln(A + ε) ≥ voln(B + ε), where B is a spherical ball with the same volume as A. THE ISOPERIMETRIC INEQUALITY ON COMPACT RANK ONE SYMMETRIC SPACES 13

Proof. Let 0 < κ1 < 1 and ε > 0 (small enough) be given such that κ1 + ε < 1. Let κ2 = 1 − (κ1 + ε). We apply Theorem 2.1 which gives us: n n Sep(S , κ1, κ2) ≤ NS (κ1, κ2).

n Let us now investigate the right-hand side of the above inequality, i.e. NS (κ1, κ2). Since we are dealing with needles upon which the density of the probability measures are sinn−1-concave functions, proposition 3.1 explicitly gives: n−1 n NS (κ1, κ2) = Sep(C sin(t) dt, κ1, κ2), where C is the normalizing constant.

n On the other hand, we know that NS (κ1, κ2) can be realized from two spherical balls B1 and B2 where µ(Bi) = κi with the balls being centered at opposite points. Indeed, the following formula holds for the measure (normalized volume) of a ball of radius r (centered at a point x ∈ Sn) on the canonical sphere: Z r sin(t)n−1dt vol (B(x, r)) n = 0 . n π/2 voln(S ) Z sin(t)n−1dt 0

We chose r > 0 such that the right hand side in the equation above is equal to κ1. The proof of the theorem follows by applying proposition 2.2.  Here, we present the Levy-Gromov isoperimetric inequality: Theorem 3.2 (Levy-Gromov). Let M be a compact Riemannian manifold of di- mension n such that Ricci(M) ≥ Ricci(Sn). We denote the Riemannian volume n n on M and on S by voln. Let A ⊂ M be an open subset of M and let B ⊂ S be a ball on the sphere such that

voln(A) voln(B) = n . voln(M) voln(S ) For every ε > 0 we have:

voln(A + ε) voln(B + ε) ≥ n . voln(M) voln(S ) Proof. The proof is similar to the proof of Theorem 3.1. Let κ1, κ2, ε be defined as in the proof of Theorem 3.1. We apply Theorem 2.1 which gives us:

Sep(M, κ1, κ2) ≤ NM (κ1, κ2). For any needle ν on M, the measure ν has density which is sinn−1-concave function. Moreover, the comparison of the diameter of M and the one of Sn shows that the length of the support of ν is at most equal to π. Hence, we can use the result of proposition 3.1 and obtain that n−1 NM (κ1, κ2) ≤ Sep(C sin(t) dt, κ1, κ2), where C is the normalizing constant. The rest follows as in the proof of Theorem 3.1 and the proof of theorem follows.  14 YASHAR MEMARIAN

Remark 3.1. The power of Theorem 2.1 lies in the fact that in order to obtain solution(s) to the isoperimetric problem, it is sufficient to find subsets which realize the optimal needle (candidate) separation distance. For the case of the canonical sphere, this becomes straightforward thanks to proposition 3.1. Other interesting examples will follow in the next sections.

4. Sharp Isoperimetric Inequalities on Compact Rank One Symmetric Spaces The isoperimetric problem, even on a highly symmetric manifold, is usually very hard to solve. For instance, the isoperimetric inequality on (real) projective spaces was only solved up to dimension 3 (see [23],[22]). By applying our Theorem 2.1, we are able to solve the isoperimetric problem for the real, complex and quaternionic projective spaces in every dimension.

n Theorem 4.1. Let M be the canonical real projective space RP . Let voln denote the canonical Riemannian volume of RP n. Let A be an open subset of RP n. For every ε > 0 we have:

voln(A + ε) ≥ voln(B + ε), where B is

• either an intrinsic ball with volume = voln(A). • or a tube around a totally geodesic RP k ⊂ RP n for a 1 ≤ k ≤ n − 1 with volume = voln(A). Theorem 4.2. Let M be the canonical complex projective space CP n. The (canon- ical) metric is the Fubini-Study metric with sectional curvature 1 ≤ Sec(M) ≤ 4. The volume element associated with the canonical Riemannian metric is denoted by vol2n. Let A be an open subset of M. For every ε > 0 we have:

vol2n(A + ε) ≥ vol2n(B + ε), where B is either

• an intrinsic ball with volume equal to vol2n(A), k n • or a tube around a CP ⊂ CP with volume equal to vol2n(A). Theorem 4.3. Let M be the canonical quaternionic projective space HP n. The (canonical) metric is the Fubini-Study metric with sectional curvature 1 ≤ Sec(M) ≤ 4. The volume element associated with the canonical Riemannian metric is denoted by vol4n. Let A be an open subset of M. For every ε > 0 we have:

vol4n(A + ε) ≥ vol4n(B + ε), where B is either

• an intrinsic ball with volume equal to vol4n(A), k n • or a tube around a HP ⊂ HP with volume equal to vol4n(A). Remark 4.1. • In [23], [22], the authors show that the isoperimetric solution on RP 3 are given by geodesic balls or tubes around geodesics. For different values of the volume, the isoperimetric solution (region) changes: for small or big volumes the solutions are the geodesic balls, while for the volumes THE ISOPERIMETRIC INEQUALITY ON COMPACT RANK ONE SYMMETRIC SPACES 15

around half of the total volume of RP 3 the solution of the isoperimetric problem is given by a tube around RP 1. • As pointed out to us by a referee, in [2], it is shown that for the complex projective space, some tubes around some CP k ⊂ CP n which have the same volume of the geodesic ball in CP n have smaller volume of their ε- neighborhood compared to the volume of ε-neighborhood of the geodesic ball. This already indicated the delicacy of the isoperimetric regions in the case of real, complex and quaternionic projective spaces. 4.1. Strategy of Proof of Theorems 4.1, 4.2 and 4.3. The strategy of the proof is to first show that for every κ1, κ2 > 0 such that

κ1 + κ2 < 1, The following holds. • For RP n, there exists a k ∈ {0, ··· , n − 1} such that: n−k−1 k n NRP (κ1, κ2) = Sep(Ck sin(t) cos(t) dt, κ1, κ2),

where Ck is the normalizing constant and the support is (0, π/2). • For CP n, there exists a k ∈ {0, ··· , n − 1} such that: 2n−2k−1 2k+1 n NCP (κ1, κ2) = Sep(Ck sin(t) cos(t) dt, κ1, κ2),

where Ck is the normalizing constant and the support is (0, π/2). • For HP n, there exists a k ∈ {0, ··· , n − 1} such that: 4n−4k−1 4k+3 n NHP (κ1, κ2) = Sep(Ck sin(t) cos(t) dt, κ1, κ2),

where Ck is the normalizing constant and the support is (0, π/2). Once this is proven, it is sufficient to apply Proposition 2.2 and find those subsets which realize the needle (candidate) separation distances. The proof relies on the specific geometry of each of the above spaces in consider- ation. Therefore, we treat each space separately and present a proof for each case. We start with the real projective space, then treat the complex and quaternionic projective spaces.

5. The proof of Theorem 4.1 for RP n The study of isoperimetric inequality on RP n using the needles is different from the case of the round sphere even though the canonical RP n is of constant sectional curvature 1. Since minimizing geodesics on RP n have length at most equal to π/2, (almost) the only (but crucial) difference for needles on RP n from the needles defined on the canonical sphere Sn is the maximum length of their support. Hence, we are restricted to study sinn−1-concave measures on intervals with maximum length π/2. This difference has a big impact on the result when studying separation dis- tance of needles in RP n. To get a feeling about this difference, we observe that on the sphere there is only one needle with maximal support which is the needle C sin(t)ndt, however on the real projective space, there are several needles of max- k n−k imal support: Ck sin(t) cos(t) dt. This fact plays a crucial role for the study of isoperimetric inequality of RP n using the needle decomposition. In order to prove Theorem 4.1, we need two propositions, one purely analytic and the other with geometrical flavor. We start with the required analytic result: 16 YASHAR MEMARIAN

Proposition 5.1. Let n > 1. Let

f(t) = C1f1(t)f2(t), where

f1(t) = cos(t + c1) ··· cos(t + cp)

f2(t) = cos(t + cp+1) ··· cos(t + cp+k).

Where C1 > 0 and p + k = n − 1.

For i ∈ {1, ··· p} we have ci ≥ 0, and for j ∈ {p + 1 ··· , p + k} we have cj < 0. We assume µ1 = f(t)dt is a probability measure on [0,L] where L ≤ π/2 and where f(L) = 0. Moreover, if f(0) 6= 0 then k = 0. Let k p g(t) = C2 sin(t) cos(t) .

We assume µ2 = g(t)dt is a probability measure on [0, π/2]. Then for every κ1, κ2 > 0 such that

κ1 + κ2 < 1, we have:

Sep(µ2, κ1, κ2) ≥ Sep(µ1, κ1, κ2). Proof. When k = 0, the proof is similar to the proof of Proposition 3.1. We assume then k 6= 0 for which case the proof is more involved. We start by the following:

Lemma 5.1. Let t1 ∈ [0,L] (resp. t2 ∈ [0, π/2]) be the point where f (resp. g) attains its maximum. Then

t1 ≤ t2. Proof. We have: p+k f(t)0 X = − tan(t + c ) f(t) i i=1 ≤ −p tan(t) + k cot(t) g(t)0 = . g(t) 0 Since we have f, g ≥ 0 and since t1 is such that f(t1) = 0 then from the above 0 inequality we deduce that at point t1, g(t1) ≥ 0, which means in the interval (0, t1), the function g is increasing. Therefore, we have:

t1 ≤ t2. This ends the proof of this Lemma. 

Let c be the maximum of {c1, ··· , cp}. We shift the function f by c to the right, i.e. the transformation t → t − c. We define then F (t) = f(t − c), and we have:

F (t) = C1 cos(t + d1) ··· cos(t + dp+k), THE ISOPERIMETRIC INEQUALITY ON COMPACT RANK ONE SYMMETRIC SPACES 17 where for every i ∈ {1, ··· , p + k} we have

di = ci − c.

Clearly, F (π/2) = 0 and the number of those di < 0 is at least equal to k. The function F is the function f shifted by c in the direction of positive axis. Since we studied the function f on the interval [0,L], we study the function F on [c, π/2]. The following Lemma is the counterpart of Lemma 5.1:

Lemma 5.2. Let t1 ∈ [c, π/2] (resp. t2 ∈ [0, π/2]) be the point where F (resp. g) attains its maximum. Then

t1 ≥ t2. Proof. The Proof of this Lemma is similar to the proof of Lemma 5.1. We have:

p+k F (t)0 X = − tan(t + d ) F (t) i i=1 ≥ −p tan(t) + k cot(t) g(t)0 = . g(t) 0 Since on the interval [c, π/2] we have F ≥ 0 and since t1 is such that F (t1) = 0 0 then from the above inequality we deduce that g(t1) ≤ 0. Therefore, on (t1, π/2) the function g is decreasing. Hence we have:

t1 ≥ t2. This ends the proof of this Lemma. 

According to Lemma 5.1, we have t1 ≤ t2. We translate the function f by t2 −t1 to the right, i.e. we perform the transformation t → t−(t2 −t1). Let the translated function be denoted by Fm. According to Lemma 5.2, we have

[t2 − t1,L + (t2 − t1)] ⊂ [0, π/2].

Moreover, the maximum points of Fm and g coincide. Recall that F dt is a proba- bility measure on [c, π/2] hence µm = Fmdt is a probability measure on

[t2 − t2,L + (t2 − t2)].

Moreover, both Fm and g are increasing from (t2 −t1, t2) and decreasing on (t2,L+ (t2 − t1)). Hence we have:

Fm(t2) ≥ g(t2).

Indeed otherwise, the function Fm would be strictly smaller than g which is not possible since both g(t)dt and Fm(t)dt are probability measures. We are set to finalize the proof of proposition 5.1. Let κ1, κ2 > 0 with

κ1 + κ2 < 1, be given. Recall µ1 = f(t)dt. It is clear that:

Sep((µ1, [0,L]), κ1, κ2) = Sep((µm, [t2 − t1,L + (t2 − t1)]), κ1, κ2). 18 YASHAR MEMARIAN

Indeed, assume the sets U1,U2 ⊂ [0,T ] where

µ1(Ui) = κi, for i = 1, 2 realize the separation distance Sep(µ1, κ1, κ2) which we know from Lemma 3.2 that are connected intervals. Then for i = 1, 2, the sets

Wi = Ui + (t2 − t1) realize the separation distance Sep(µm, κ1, κ2). For t ∈ (t2 − t1,L + (t2 − t1)), consider the function:

q(t) = µ2((0, t)) − µm((t2 − t1, t)). It is clear that q is a continuous function and it is clear that for t small enough we have q(t) > 0, and for t large enough we have: q(t) < 0.

Indeed, for t small enough we have g(t) ≥ Fm(t) hence µ2(0, t) ≥ µm(t2 − t1, t). And for t = L + (t2 − t1), we have µm((t2 − t1, t)) = 1 and µ2((0, t)) < 1. Therefore, there exists a t0 ∈ (t2 − t1,L + (t2 − t1)) such that:

q(t0) = 0. The above translates to the fact that:

µm(t2 − t1, t0) = µ2(0, t0) = κ, for some κ ∈ (0, 1). We have different possibilities for the values of κ1 and κ2 summarized as follows and treated separately:

• We have κ1 ≤ κ and κ2 ≤ 1 − κ. • We have κ1 ≤ κ and κ2 ≥ 1 − κ. • We have κ1 ≥ κ and κ2 ≤ 1 − κ.

For the first case, where κ1 ≤ κ and κ2 ≤ 1−κ, by the properties of the function q we have:

µm(W1) ≤ µ2(W1)

µm(W2) ≤ µ2(W2).

Therefore, for F1,F2 ⊂ [0, π/2] such that:

µ2(F1) = κ1

µ2(F2) = κ2, we have

d(F1,F2) ≥ d(W1,W2), where we refer to definition 2.2 for the definition of the distance d(., .) between the intervals. We now treat the second case, where κ1 ≤ κ and κ2 ≥ 1 − κ. Let F1 = (0, u1) ⊂ (0, π/2) be the interval for which we have

µ2(F1) = κ1. THE ISOPERIMETRIC INEQUALITY ON COMPACT RANK ONE SYMMETRIC SPACES 19

Let W1 = (t2 − t1, w1) such that:

µm(W1) = κ1.

Since κ1 ≤ κ, we have:

u1 ≤ w1. Let

W2 = (w2,L + (t2 − t1))

F2 = (u2, π/2), such that, we have:

µ2(F2) = µm(W2)

= κ2. Suppose the following inequality holds:

(5.1) |u1 − w1| ≤ |u2 − w2|. It is clear that we have:

(5.2) µ2(u1, u2) = µm(w1, w2).

From inequality (5.1), we assumed the length of the interval (u1, u2) be larger than the length of the interval (w1, w2). Further, the function Fm is larger than g on the interval (w1, w2) which contradicts equation (5.2). Therefore, we have:

d(F1,F2) ≥ d(W1,W2).

We treat now the last case for κ1, κ2 where we have κ1 ≥ κ and κ2 ≤ 1 − κ. The proof is similar to the previous case. Let F1 = (0, u1) ⊂ (0, π/2) be the interval for which we have

µ2(F1) = κ1.

Let W1 = (t2 − t1, w1) such that:

µm(W1) = κ1.

Since κ1 ≥ κ, we have:

u1 ≥ w1. Let

W2 = (t2 − t1, w2)

F2 = (0, u2), such that we have:

µ2(F2) = µm(W2)

= 1 − κ2. Suppose we have:

|u1 − w1| ≥ |u2 − w2|, It is clear that we have:

(5.3) µ2(u1, u2) = µm(w1, w2). 20 YASHAR MEMARIAN

We assumed the length of the interval (u1, u2) be larger than the length of the interval (w1, w2). Further, the function Fm is larger than g on the interval (w1, w2). Therefore, it contradicts equation (5.3). Therefore, we have:

d(F1,F2) ≥ d(W1,W2).

We have demonstrated that for every κ1, κ2 > 0 such that

κ1 + κ2 < 1, we have

(5.4) d(F1,F2) ≥ d(W1,W2). According to Lemma 3.2 we know that the separation distance is realized by con- nected intervals, and using the inequality (5.4) we deduce that:

Sep(µ2, κ1, κ2) ≥ d(F1,F2)

≥ d(W1,W2)

= Sep(µm, κ1, κ2). The proof of Proposition 5.1 is completed.  Remark 5.1. In the proof of Proposition 5.1, we showed that the support of the measure µm is strictly included in the support of the measure µ2 which is the interval (0, π/2). Then it was easy to demonstrate that the separation distance of µ2 is larger than the separation distance of the measure µm. The smaller the support of a measure is, the larger the maximum point would be and the measure is more concentrated around its maximum point which makes the separation distance less than a measure with larger support. This idea was already known and is explained in [13] (pages 154 − 155). Having proved the analytic result of proposition (5.1), we present a result which has to do with the geometry of RP n: Proposition 5.2. Let n ≥ 1. Let ν be a needle candidate on RP n. Let f be the density of the pull-back of ν on A = (0,L) with L ≤ π/2. Then

n−1 f(t) = C Π1 fi(t), where C ∈ R and for every i ∈ {1, ··· , n − 1} there exists βi ∈ R such that

fi(t) = sin(t + βi).

Proof. Let the dimension n ≥ 1 of RP n be fixed. Based on definition 2.4 of the needle candidates we know there exists a family of Jacobi fields J1(t), ··· Jn−1(t) 0 along γ(t) for t ∈ A. As in the definition 2.4, we let Jn = γ . We assume this family consists of linearly independent (tangent) vectors. Otherwise there is nothing to prove. There exists a constant c > 0 and the density of the pull-back of the needle ν, denoted by f, equals: q f(t) = c det (< Ji(t),Jk(t) >)i,k=1,··· ,n. THE ISOPERIMETRIC INEQUALITY ON COMPACT RANK ONE SYMMETRIC SPACES 21

n The Riemannian volume voln of RP can be transported by the (inverse of the) exponential map on the tangent space at any point. By Linear algebra, we know that q det (< Ji(t),Jk(t) >)i,k=1,··· ,n,

n is the determinant of the Gram matrix associated to vectors Ji(t) ∈ Tγ(t)RP for i = 1, ··· , n. This quantity is equal to the (Riemannian) volume of the parallelepiped n generated by the linearly independent tangent vectors {Ji(t)}i=1 in the tangent n space Tγ(t)RP . We denote by:

[J1, ··· ,Jn], the parallelepiped, generated by the vectors Ji (i = 1, ··· , n). For every t ∈ A, we n denote (abusively) by voln the Riemannian volume in the tangent space Tγ(t)RP . By definition 2.4, there exists a point t ∈ A on which we have: 0 0 0 < Ji(t), γ (t) > = < Ji (t), γ (t) > = 0 for i = 1, ··· , n − 1 and 0 0 < Ji (t),Jk(t) > = < Jk(t),Ji(t) >, for i, k = 1, ··· , n − 1. An easy exercise concerning the Jacobi fields (see [24],[7]) asserts that if such equality holds for one point, then it holds on every point on γ(t). 0 Thus, the frame of Jacobi fields Ji are orthogonal to γ for every t ∈ A. Further- more, without loss of generality we may assume that {Ji(t)} form an orthogonal basis of the tangent space at γ(t). Indeed, since for i = 1, ··· , n − 1, the vectors {Ji(t)} are linearly independent, there exists a constant (n − 1) × (n − 1) matrix A (along γ(t) such that {AJi(t)} are linearly independent and orthogonal Jacobi fields. We have the following result in : Proposition 5.3. Let M be a Riemannian manifold of dimension n of constant sectional curvature 1. Let σ be a geodesic segment in M. Let {E1, ··· ,En} be orthogonal and parallel along σ with E1 be the unit tangent vector along σ. Then the general solution of a Jacobi field along σ is given by: X (ai sin(t) + bi cos(t))Ei. i

Therefore, by applying proposition 5.3 to the orthogonal Jacobi fields {AJi(t)} the proof follows.  5.1. End Proof of Theorem 4.1. We quickly remind some known facts on Rie- mannian geometry: Definition 5.1. Let M be a closed Riemannian manifold of dimension n and let H ⊂ M be a submanifold of dimension 0 < k < n. Let c :[a, b] → M be a geodesic in M such that: • The geodesic c is orthogonal to H. • c(a) ∈ H. 0 ⊥ • c(a) ∈ Tc(a)H . 22 YASHAR MEMARIAN

Then, c(t0) ∈ c(a, b] is called a focal point of H if there exists a non-trivial Jacobi field J along c with J(a) ∈ Tc(a)H such that J(t0) = 0. Furthermore, the distance between c(a) to c(t0) is called the focal distance of c(t0) to H. We invite the reader to consult [16] and [17] for some related topics on focal distances of submanifolds in RP n, CP n or HP n. For our purpose, we need to know that the maximal focal distance of any submanifold in RP n is equal to π/2. Therefore, according to Lemma 3.2 and particularly inequality (3.2), for the purpose of estimating the separation distance, it is enough to consider those needle candidates ν for which the density of their pull-back is defined on an interval (0,L) with L ≤ π/2 and is equal to f(t) as it is defined in Proposition 5.1. Furthermore, according to this same proposition 5.1 we deduce that needle candidates of the form

g(t) = sin(t)m1 cos(t)m2 , with m1, m2 ≥ 0 and m1 + m2 = n − 1 maximize the needle separation distance. Therefore, for κ1, κ2 > 0 such that

κ1 + κ2 < 1, there exists a k ∈ {1 ··· , n} such that:

n−k k−1 n NRP (κ1, κ2) = Sep(Ck sin(t) cos(t) dt, κ1, κ2), Therefore, according to Proposition 2.2, it remains to find those subsets of RP n which realize the above (optimal) needle separation distance. The measure (normalized volume) of balls in RP n is the same as the measure of balls in Sn. Indeed, for 0 < r ≤ π/2, and x ∈ RP n: Z r sin(t)n−1dt vol (B(x, r)) n = 0 . n π/2 voln(RP ) Z sin(t)n−1dt 0 The measure (normalized volume) of a tube around a totally geodesic

n−k n RP ⊂ RP is given by: Z r sin(t)k−1 cos(t)n−kdt vol ( P n−k + r) n R = 0 . n π/2 voln(RP ) Z sin(t)k−1 cos(t)n−kdt 0 One could consult [10] for a proof of the above formula. We apply the result of Proposition 2.2 and Lemma 5.1 and deduce that for every open set U ⊂ RP n we have:

voln(U + δ) ≥ voln(B + δ), where B is either a ball or a tube around some RP k where:

voln(U) = voln(B). This ends the proof of Theorem 4.1. THE ISOPERIMETRIC INEQUALITY ON COMPACT RANK ONE SYMMETRIC SPACES 23

6. The proof of Theorem 4.2 for CP n n We inherit CP with the canonical Riemannian structure and denote vol2n by the Riemannian volume. We start with an analytic result:

Proposition 6.1. Let n ≥ 1. Let

f(t) = C1f1(t)f2(t)f3(t), where

f1(t) = cos(t + a1) ··· cos(t + ap)

f2(t) = cos(t + b1) ··· cos(t + bq)

f3(t) = sin(2t + c1) ··· sin(2t + cs). where C1 > 0, s ≥ 1 and p + q + s = 2n − 1.

Moreover for i ∈ {1, ··· max{p, q, s}} we have ai, ci ≥ 0 and bi < 0. We assume µ1 = f(t)dt is a probability measure on [0,L] where L ≤ π/2 and where f(0) = f(L) = 0.

Let

m1 m2 m3 g(t) = C2 sin(t) cos(t) sin(2t) , where

m1 + m2 + m3 = 2n − 1, and

m2 ≤ p

q ≤ m1

s ≤ m3.

Moreover we assume µ2 = g(t)dt is a probability measure on [0, π/2]. Then for every κ1, κ2 > 0 where

κ1 + κ2 < 1, we have:

Sep(µ2, κ1, κ2) ≥ Sep(µ1, κ1, κ2). Proof. The strategy of proof is similar to the proof of proposition 5.1. We start by the following:

Lemma 6.1. Let t1 ∈ [0,L] (resp. t2 ∈ [0, π/2]) be the point where f (resp. g) attains its maximum. Then

t1 ≤ t2. 24 YASHAR MEMARIAN

Proof. On the interval [0,L] we have: p q s f(t)0 X X X = − tan(t + a ) − tan(t + b ) + 2 cot(t + c ) f(t) i i i i=1 i=1 i=1

≤ −m2 tan(t) + m1 cot(t) + m3 cot(2t) g(t)0 = . g(t) 0 Since we have f, g ≥ 0 and since t1 is such that f(t1) = 0 then from the above 0 inequality we deduce that at point t1, g(t1) ≥ 0. Hence the function g is increasing on the interval (0, t1). Therefore, we have:

t1 ≤ t2. This ends the proof of this Lemma. 

c1 cs Let c be the maximum of {a1, ··· , ap, 2 ··· , 2 }. We shift the function f by c i.e. the transformation t → t − c. We define then F (t) = f(t − c) and we have:

F (t) = C1f1(t − c)f2(t − c)f3(t − c), and we observe that F (π/2) = 0. The function F is the function f shifted by c in the direction of positive axis. Since we studied the function f on the interval [0,L], we study the function F on [c, π/2]. The following Lemma is the counterpart of Lemma 6.1:

Lemma 6.2. Let t1 ∈ [c, π/2] (resp. t2 ∈ [0, π/2]) be the point where F (resp. g) attains its maximum. Then

t1 ≥ t2. Proof. We make the change of variable t → π/2 − t and observe that:

m1 m2 m3 g(π/2 − t) = C2 cos(t) sin(t) sin(2t) , and F (π/2 − t) = f(π/2 − t − c)

= C1f1(π/2 − t − c)f2(π/2 − t − c)f3(π/2 − t − c), where

f1(π/2 − t − c) = cos(π/2 − t + a1 − c) ··· cos(π/2 − t + ap − c)

f2(π/2 − t − c) = cos(π/2 − t + b1 − c) ··· cos(π/2 − t + bq − c)

f3(π/2 − t − c) = sin(2t + c1 − 2c) ··· sin(2t + cs − 2c).

Since ai, ci ≥ 0 and bi < 0 and c is the maximum of {a1, ··· , ap, c1 ··· , cs}, we have:

f1(π/2 − t − c) = cos(t + d1) ··· cos(t + dp)

f2(π/2 − t − c) = cos(t + h1) ··· cos(t + hq)

f3(π/2 − t − c) = sin(2t + c1 − 2c) ··· sin(2t + cs − 2c), where di ≤ 0 and hi ≥ 0. Therefore, we apply Lemma 6.1 to F (π/2 − t) and g(π/2 − t) and the proof follows.  THE ISOPERIMETRIC INEQUALITY ON COMPACT RANK ONE SYMMETRIC SPACES 25

According to Lemmas (6.1) and (6.2), the function f(t − (t2 − t1)) obtains its maximum at the point t2 which coincides with the point where g attains its maxi- mum. Moreover, we have:

[t2 − t1,L + (t2 − t1)] ⊂ [0, π/2].

Hence, the support of f(t − (t2 − t1)) is strictly contained in the support of g and therefore:

f(t2) ≥ g(t1). The rest of the proof is exactly similar to the proof of proposition (5.1). Therefore:

Sep(µ2, κ1, κ2) ≥ Sep(µ1, κ1, κ2). 

Remark 6.1. In Proposition 6.1 we could have replaced the function f1(t)f2(t)f3(t) by suitable functions of the form: p Πi=1 sin(αit + βi), where αi ∈ [1, 2] and βi ∈ R such that f2(0) = f2(L) = 0. The proof is indeed similar.

Next result concerns the needle candidates on CP n and the result will be decisive for the proof of the isoperimetric inequality on CP n. Proposition 6.2. Let n ≥ 1. Let ν be a needle candidate on CP n. Let f be the density function of the pull-back of ν on an interval A = (0,L) ⊂ R such that L ≤ π/2. Then 2n−1 f(t) = Πi=1 fi(t), where for every i ∈ {1, ··· , 2n − 1} there exists αi ∈ [1, 2] and βi ∈ R such that

fi(t) = sin(αit + βi). Proof. Let ν be a needle candidate supported on a geodesic γ ⊂ CP n which is parametrized by its arclength t ∈ A ⊆ [0, π/2]. Let f be the density function of the pull-back of ν. All we know about f at this stage is that f is a sin2n−1-concave function. We will dig more and find more properties for f which has to do with the geometry of CP n. Lemma 6.3. Let the interval (0,L) be as above. Let f be the density of the pull-back of the needle candidate ν. Then

f = f1f2 where f1 is a function satisfying the inequality: 00 f1 + 4f1 ≤ 0, 2n−2 and f2 is a sin -concave function. Proof. The complex projective space has real dimension 2n. Based on defini- tion 2.4 of the needle candidates we know there exists a family of Jacobi fields 0 J1(t), ··· J2n−1(t) along γ for t ∈ A. Let J2n = γ and assume this family consists of linearly independent (tangent) vectors. 26 YASHAR MEMARIAN

There exists a constant c > 0 and the density of the needle ν, denoted by f, equals: q f(t) = c det (< Ji(t),Jk(t) >)i,k=1,··· ,2n.

n The Riemannian volume vol2n of CP can be transported by the (inverse of the) exponential map on the tangent space at any point. By Linear algebra, we know that q det (< Ji(t),Jk(t) >)i,k=1,··· ,2n,

n is the determinant of the Gram matrix associated to vectors Ji(t) ∈ Tγ(t)CP for i = 1, ··· , 2n. This quantity is equal to the (Riemannian) volume of the parallelepiped 2n generated by the linearly independent tangent vectors {Ji(t)}i=1 in the tangent n space Tγ(t)CP . We denote by:

[J1, ··· ,J2n], the parallelepiped, generated by the vectors Ji (i = 1, ··· , 2n). For every t ∈ A we denote (abusively) by vol2n the Riemannian volume in the tangent space n Tγ(t)CP . Here, the space of Jacobi fields along a geodesic γ is a linear vector space of dimension 4n. By definition 2.4, there exists a point t ∈ A on which we have: 0 0 0 < Ji(t), γ (t) > = < Ji (t), γ (t) > = 0 for i = 1, ··· , 2n − 1 and 0 0 < Ji (t),Jk(t) > = < Jk(t),Ji(t) >, for i, k = 1, ··· , 2n − 1. An easy exercise concerning the Jacobi fields (see [24],[7]) asserts that if such equality holds for one point, then it holds on every point on 0 γ(t). Thus, the frame of Jacobi fields Ji are orthogonal to γ for every t ∈ A. For every i = 1, ··· , 2n, let

Ji(t) = fi(t)Ei(t), where kEi(t)k = 1. With the same argument presented in the proof of proposition (5.2) we assume the family {Ei(t)} forming an orthonormal basis of the tangent space at the point γ(t) (and hence on every point along γ(t)). n Let t ∈ A, and let γ(t) be a point on the geodesic γ. Let Tγ(t)CP be the tangent space at this point. By linear algebra we know that:

vol2n([J1(t), ··· ,J2n(t)]) = kJ2n(t)kvol2n−1([J1(t), ··· ,J2n−1])

= vol2n−1([J1(t), ··· ,J2n−1]). Indeed, the geodesic γ is parametrized by its arclength, hence 0 kJ2nk = kγ k = 1. Moreover,

[J1(t), ··· ,J2n−1], is the parallelepiped generated by Ji(t) for i = 1, ··· , 2n − 1 and is contained in n 0 a (2n − 1)-dimensional vector subspace of Tγ(t)CP (which is orthogonal γ (t)) . THE ISOPERIMETRIC INEQUALITY ON COMPACT RANK ONE SYMMETRIC SPACES 27

2n This parallelepiped is in fact a (2n − 1)-dimensional rectangle as {Ji(t)}i=1 form n an orthogonal basis of the tangent space Tγ(t)CP . At the point γ(t), consider the orthonormal basis:

e1, e2, ··· , e2n, where for a m ∈ {1, ··· , 2n}, we have:

em = iem−1, where iem−1 is the complex multiplication of em−1 by the complex number i. There- 2 n fore, the vectors em−1 and em form an orthonormal base of C ⊂ Tγ(t)CP . Again, by linear algebra, we have:

vol2n−1([J1(t), ··· J2n−1(t)] = kJm(t)kvol2n−2(R2n−2(t)), where

◦ R2n−2(t) = [J2(t), ··· ,Jm(t) , ··· ,J2n−1(t)], is the (2n − 2)-dimensional rectangle on which Jm is orthogonal. n By the property of CP , the sectional curvature of the 2-plane {em, iem} is equal n to 4. Since CP is a , for every t ∈ A the Jacobi field Jm, remains parallel along γ(t) . This is a strong property which relies on the symmetry of CP n and can be verified in more detail in [7] and [24]. 2 Hence, by definition, Jm is a Jacobi field on a S (4) which is a sphere of constant curvature 4 and thus,

f1(t) = kJm(t)k, satisfies the following differential inequality:

00 (6.1) f1 + 4f1 ≤ 0. Let

◦ f2(t) = vol2n−2([J2(t), ··· ,Jm, ··· ,J2n−1(t)]).

Since the norm of the Jacobi fields Ji(t) for every i ∈ {2, ··· , 2n − 2} is a sin- concave function for t ∈ A, therefore according to lemma 3.2, the function f2(t) is a sin2n−2-concave function. This ends the proof of the lemma.  Let

2n+1 n p : S → CP be the Riemannian submersion from the canonical 2n+1-dimensional sphere to the n-dimensional complex projective space. Let γ ⊂ CP n be a (minimal) geodesic on the complex projective space. It is known (see for instance [24]) that for every Jacobi field J along γ , there exists a geodesicγ ˆ on S2n+1 and a Jacobi field along this geodesic which projects on γ and J. According to Proposition (5.3) we know the general form of orthogonal Jacobi fields along a geodesic on S2n+1. Therefore, by projection on CP n and applying lemma 6.3 the proof of the proposition is completed.  28 YASHAR MEMARIAN

For our purpose, we could go further and only consider those needle candidates ν such that the density of their pull-back is given by:

C f1(t)f2(t), where C ∈ R and where p f1(t) = Πi=1 sin(t + αi) k f2(t) = Πi=1 sin(2t + βi).

Where αi, βi ∈ R and k + p = 2n − 1, with k ≤ p. Indeed we have:

Lemma 6.4. Let ν be a needle candidate defined on CP n. Let the support of ν be a geodesic γ parametrized by the interval (0,L). Let U ⊂ CP n be a domain such that for every γ(t), the volume of the (2n − 1)-dimensional hyper-surface Ut ⊂ U which 0 is orthogonal to γ (t) and which is generated by {J1(t), ··· ,J2n−1(t)} is equal to q f(t) = c det (< Ji(t),Jk(t) >)i,k=1,··· ,2n−1.

Then, there exists a frame of orthogonal Jacobi fields {Q1(t), ··· ,Q2n−1(t)} along γ such that the (2n − 1)-dimensional volume of Ut is equal to:

Cf1(t)f2(t), for some C ∈ R and f1, f2 defined as above.

Proof. We consider the exponential map with respect to the submanifold N = U0. For this we need to define the normal bundle ν(N) which by definition is all the pairs (x, v) such that x ∈ N and v is orthogonal to TxN. Therefore the exponential map: n expN : ν(N) → CP , is defined by

expN (x, v) = expx(v). For more details on exponential maps with respect to submanifolds we refer to [10]. 0 n Let p = γ(0) ∈ N where γ(0) is orthogonal to TpN. In TpCP if we consider the straight line t → tγ(0)0 and along this line we consider a linear vector field X(t) = a + tb where a, b ∈ TpN are orthogonal vectors which are moved parallel to themselves, then the map dexpN sends this field to normal Jacobi fields along γ. It is a classical result and can be verified in textbooks that all orthogonal jacobi fields along γ are obtained in this way. Therefore it is sufficient to chose suitable orthogonal linearly independent Xi = ai + bit and take Yi = dexpN Xi such that: 2n−1 Πi=1 kYik = Cf1(t)f2(t). This ends the proof of the Lemma.  Proposition 6.1 suggests that the needle separation distance (on CP n) is obtained 2n−k k by the separation distance of the needles of the form µk = Ck sin(t) cos(t) dt for some k ∈ {1, ··· , 2n − 1}. This is already helpful and leaves one to find those (open) subsets of CP n which realize the separation distances of the needles of form µk . The next proposition classifies this problem: THE ISOPERIMETRIC INEQUALITY ON COMPACT RANK ONE SYMMETRIC SPACES 29

2n−k k Proposition 6.3. Let n ≥ 1. Let µk = Ck sin(t) cos(t) dt for k ∈ {1 ··· , 2n − 1}, be a probability measure on [0, π/2]. Let κ1, κ2 > 0 be such that:

κ1 + κ2 < 1. n Then, there exist U1,U2 ⊂ CP with

vol2n(Ui) n = κi, vol2n(CP ) for i = 1, 2 such that:

d(U1,U2) = Sep(µk, κ1, κ2), if and only if we have k = 2m + 1 for m ∈ {0, ··· n − 1}. Proof. If k = 2m − 1 for 1 ≤ k ≤ n − 1, the subset which realizes the separation distance of µk is either a geodesic ball (and its complementary) or a tube around CP k (and its complementary). Indeed, the formula for the volume of a (geodesic) ball in CP n can be found in classic Riemannian Geometry textbooks such as in [7] and [10]. Let x ∈ CP n, let r > 0, the ball B(x, r) ⊂ CP n is by definition all the points within distance δ to the point x. The measure (normalized volume) is given by: Z r sin(t)2n−1 cos(t)dt vol (B(x, r)) 2n = 0 . n π/2 vol2n(CP ) Z sin(t)2n−1 cos(t)dt 0 And let k ≥ 1 and let CP k ⊂ CP n be a totally geodesic complex submanifold of CP n. Let δ > 0. Let CP k + δ be a tube of radius δ around this CP k. The formula for the volume of such tubes can be found in [10] and are given by: Z r sin(t)2n−2k+1 cos(t)2k−1dt vol ( P k + r) 2n C = 0 . n π/2 vol2n(CP ) Z sin(t)2n−2k+1 cos(t)2k−1dt 0 n Now suppose k = 2m and there exist U1,U2 ⊂ CP such that for κ1, κ2 > 0 with κ1 + κ2 < 1, we have

vol2n(Ui) n = κi, vol2n(CP ) where i = 1, 2. Moreover, we have:

d(U1,U2) = Sep(µk, κ1, κ2). Let n σ : [0, π/2] → CP , be a geodesic of (maximal) length π/2 parametrized by its arc length t. Let {J1(t), ··· ,J2n(t)} be an orthogonal family of Jacobi fields along σ such that J2n(t) 30 YASHAR MEMARIAN coincides with the tangent vector along σ. We assume for every t ∈ [0, π/2]:

q 2n det(< Ji(t),Jk(t) >)i,k=1,··· ,2n = Πi=1kJi(t)k 2n−2k 2k = Ck sin(t) cos(t) .

Let S = {Jj1(t), ··· ,Jjl(t)} be the maximal subset of {J1(t), ··· ,J2n−1(t)} such that Jjβ(0) 6= 0 for every β ∈ {1, ··· , l}. Because of the choice of Jacobi fields, we have l ≥ 1. Indeed, if for every i ∈ {1, ··· , 2n − 1}, we have Ji(0) = 0, then n U1 represents a geodesic ball in CP and the needle candidate which realizes the separation distance of a geodesic ball (and its complementary) in CP n is µ = C sin(t)2n−1 cos(t)dt on a maximal geodesic parametrized by (0, π/2). By definition of Jacobi fields, for every Jjβ ∈ S there exists a (geodesic) variation: n α : [0, π/2] × (−ε, +ε) → CP , such that α is smooth and α(t, 0) = γ(t), and for every s ∈ (−ε, +ε) the curve αs(t) = α(s, t) is a geodesic and such that: ∂α (0, t) = J (t). ∂t jβ

Moreover, by (covariant) differentiating the Jacobi fields Jjβ we have 0 Jjβ(0) = 0. Hence, the image of the variation for time s = 0 lies in a totally geodesic submanifold n of CP . Therefore, the subset U1 is a tube around a totally geodesic submanifold of CP n, i.e. we have a l-dimensional totally geodesic submanifold of CP n such that U1 = N + δ for some δ > 0 and such that: Z δ sin(t)2n−2k cos(t)2kdt vol (N + δ) 2n = 0 . n π/2 vol2n(CP ) Z sin(t)2n−2k cos(t)2kdt 0

Since every Jjβ ∈ S is such that the first zero is obtained at distance π/2, i.e. Jjβ(π/2) = 0 and for every t < π/2, we have Jjβ(t) 6= 0, therefore, the totally geodesic submanifold N must be either a point or a CP k ⊂ CP n. Indeed, for totally geodesic submanifold of CP n which consist of real projective spaces RP l, the first zero of such Jacobi fields are obtained at distance π/4. This completes the proof.  6.1. End proof of Theorem 4.2. Since the maximal focal distance of any sub- manifold in CP n is equal to π/2, by applying Lemmas 3.2, 6.4 and 6.3, for the purpose of estimating the needle separation distance, we can consider only those needle candidates ν such that the density f of their pull-back is defined on the interval (0,L) where L ≤ π/2 and f is of the form presented in proposition 6.1. Hence, let µ1 = f(t)dt be a probability measure on an interval (0,L) where L ≤ π/2 where:

f(t) = C1f1(t)f2(t)f3(t), THE ISOPERIMETRIC INEQUALITY ON COMPACT RANK ONE SYMMETRIC SPACES 31 where

f1(t) = cos(t + a1) ··· cos(t + ap)

f2(t) = cos(t + b1) ··· cos(t + bq)

f3(t) = sin(2t + c1) ··· sin(2t + cs). where C1 > 0, s ≥ 1 and p + q + s = 2n − 1.

For i ∈ {1, ··· max{p, q, s}} we have ai, ci ≥ 0 and bi < 0. Moreover we assume f(0) = f(L) = 0 which is a condition which is verified according to Lemmas 6.3 and 3.2. Then, we chose m1, m2, m3 such that

m1 + m2 + m3 = 2n − 1, such that

m2 ≤ p

q ≤ m1

s ≤ m3, and

m1 m2 m3 g(t) = C2 sin(t) cos(t) sin(2t) 2n−2k+1 2k−1 = C2 sin(t) cos(t) , for some k ∈ {1 ··· , n} and some C2 > 0 such that µ2 = g(t)dt is a probability measure on (0, π/2). Therefore by applying proposition 6.1 we deduce that for κ1, κ2 > 0 such that

κ1 + κ2 < 1, there exists a k ∈ {1 ··· , n} such that: 2n−2k+1 2k−1 n NCP (κ1, κ2) = Sep(Ck sin(t) cos(t) dt, κ1, κ2), for some k ∈ {1, ··· , n}. Therefore, according to proposition 2.2 we are left to find those subsets of CP n which realize the above needle separation distances. This is presented in proposition 6.3. Hence, for every open set U ⊂ CP n we have:

vol2n(U + δ) ≥ vol2n(B + δ), where B is a (geodesic) ball or a tube around some CP k ⊂ CP n where:

vol2n(U) = vol2n(B). This ends the proof of Theorem 4.2 for the complex projective spaces.

7. The proof of Theorem 4.3 for HP n The proof for the quaternionic projective space is similar to the proof for the complex projective space. Despite this resemblance, we prefer to present the proof in complete details as the result is of great importance. Quaternionic projective space has a very rich and interesting geometry. It is defined as the quotient of the (4n+3)-dimensional sphere by the action of the group Sp(1) which is the group of unit quaternions which is isomorphism to the group structure on the 3-dimensional sphere S3. It is a real 4n-dimensional manifold. 32 YASHAR MEMARIAN

We inherit HP n with the canonical Riemannian structure (the one which makes 4n+3 n the submersion S → HP into a Riemannian submersion) and denote vol4n by the Riemannian volume associated to the canonical Riemannian structure. The sectional curvature of HP n, similar to CP n, varies between 1 and 4. Next, we continue with some geometric results.

Proposition 7.1. Let n ≥ 1. Let ν be a needle candidate on HP n. Let f be the density function of the pull-back of ν on an interval A = (0,L) ⊂ R such that L ≤ π/2. Then 4n−1 f(t) = Πi=1 fi(t), where for every i ∈ {1, ··· , 4n − 1} there exists αi ∈ [1, 2] and βi ∈ R such that

fi(t) = sin(αit + βi). Proof. Let ν be a needle candidate supported on a geodesic γ ⊂ HP n which is parametrized by its arclength t ∈ A ⊆ [0, π/2]. Let f be the density function of ν. We know that the function f is a sin4n−1-concave function. Similar to the previous section, we are interested in finding more properties for f by considering the special geometry of HP n. Lemma 7.1. Let the function f be as above. Then

f = f1f2 1 3 where f1 is a function such that q(t) = f1 satisfies the inequality: q00 + 4q ≤ 0, 4n−4 and f2 is a sin -concave function. Proof. The proof of this Lemma is similar to the proof of Lemma 6.3. The quaternionic projective space has real dimension 4n. Based on defini- tion 2.4 of the needle candidates we know there exists a family of Jacobi fields 0 J1(t), ··· J4n−1(t) along γ for t ∈ A. As in the definition 2.4, we let J4n = γ . We assume this family consists of linearly independent (tangent) vectors. Otherwise, there is nothing to prove. There exists a constant c > 0 and the density of the pull-back of the needle ν, denoted by f, equals: q f(t) = c det (< Ji(t),Jk(t) >)i,k=1,··· ,4n.

n The Riemannian volume vol4n of HP can be transported by the (inverse of the) exponential map on the tangent space at any point. We know that q det (< Ji(t),Jk(t) >)i,k=1,··· ,4n,

n is the determinant of the Gram matrix associated to vectors Ji(t) ∈ Tγ(t)HP for i = 1, ··· , 4n. This quantity is equal to the (Riemannian) volume of the paral- 4n lelepiped generated by the linearly independent tangent vectors {Ji(t)}i=1 in the n tangent space Tγ(t)HP . We denote by:

[J1, ··· ,J4n], the parallelepiped, generated by the vectors Ji (i = 1, ··· , 4n). For every t ∈ A, we denote (abusively) by vol4n the Riemannian volume in the tangent space THE ISOPERIMETRIC INEQUALITY ON COMPACT RANK ONE SYMMETRIC SPACES 33

n Tγ(t)HP . Here, the space of Jacobi fields along a geodesic γ is a linear vector space of dimension 8n. With the same argument as in the proof of Lemma 6.3 we consider the frame 0 of Jacobi fields Ji to be orthogonal to γ for every t ∈ A and the family {Ji(t)} to form an orthogonal basis of the tangent space at each point on γ(t). n Let t ∈ A and let γ(t) be a point on the geodesic γ. Let Tγ(t)HP be the tangent space at this point. We have:

vol4n([J1(t), ··· ,J4n(t)]) = kJ4n(t)kvol4n−1([J1(t), ··· ,J4n−1])

= vol4n−1([J1(t), ··· ,J4n−1]). Indeed, the geodesic γ is parametrized by its arclength, hence 0 kJ4nk = kγ k = 1. Moreover,

[J1(t), ··· ,J4n−1], is the parallelepiped generated by Ji(t) for i = 1, ··· , 4n − 1, and is contained in a n (4n − 1)-dimensional vector subspace of Tγ(t)HP . This parallelepiped is in fact a 4n (4n−1)-dimensional rectangle as {Ji(t)}i=1 form an orthogonal basis of the tangent n space Tγ(t)HP . At the point γ(t), consider the orthonormal basis:

e1, e2 ··· e4n, where there exists an m ∈ {1, ··· , 4n − 1} such that

em = iem−1

em+1 = jem−1

em+2 = kem−1, where {i, j, k} are the unit quaternions and {em−1, em, em+1, em+2} form an or- n thonormal basis of the (one-dimensional) quaternionic vector space ⊂ Tγ(t)HP . Denote this vector sub-space by T4. We have:

vol4n−1([J1(t), ··· J2n−1(t)] = vol3([Jm,Jm+1,Jm+2])vol4n−4(P4n−4), where ◦ ◦ ◦ P4n−4 = [J2(t), ··· ,Jm(t) ,Jm+1(t) ,Jm+2(t) , ··· ,J4n−1(t)], is the (4n − 4)-dimensional rectangle on which Jm,Jm+1,Jm+2 are orthogonal. By the property of HP n and with similar arguments presented in the proof of 4 Lemma 6.3, we conclude that {Jm,Jm+1,Jm+2} are Jacobi fields on a S (4) which is a 4-dimensional sphere of constant curvature 4. Thus, let

f1(t) = vol3([Jp(t),Jq(t),Js(t)]), and 1/3 q(t) = f1(t) . Then, q(t) satisfies the following differential inequality: (7.1) q00 + 4q ≤ 0. 34 YASHAR MEMARIAN

Let

f2(t) = vol4n−4(P4n−4).

Since the norm of the Jacobi fields Ji(t) for every i ∈ {2, ··· , 4n − 1} except Jm,Jm+1,Jm+2 is a sin-concave function for t ∈ A, therefore, according to lemma 4n−4 3.2, the function f2(t) is a sin -concave function. This ends the proof of the lemma.  Let 4n+3 n p : S → HP be the Riemannian submersion from the canonical 4n+3-dimensional sphere to the n-dimensional quaternionic projective space. Let γ ⊂ HP n be a (minimal) geodesic on the quaternionic projective space. Let J be a Jacobi field along γ. There exists a geodesicγ ˆ on S4n+3 and a Jacobi field along this geodesic which projects on γ and J. According to Proposition (5.3) we know the general form of orthogonal Jacobi fields along a geodesic on S4n+3. Therefore, by projection on HP n and applying lemma 7.1 the proof of the proposition is completed.  With the same arguments presented for CP n, here we can also only consider those needle candidates ν such that the density of their pull-back is given by:

C f1(t)f2(t), where C ∈ R and where p f1(t) = Πi=1 sin(t + αi) k f2(t) = Πi=1 sin(2t + βi).

Where αi, βi ∈ R and k + p = 4n − 1, with 3 ≤ k ≤ p. We now like to classify the needle candidates on HP n which have density of their pull-backs given by sin(t)m cos(t)s for some s, m with s + m = 4n + 2. This is provided by the next:

4n−k+2 k Proposition 7.2. Let n ≥ 1. Let µk = Ck sin(t) cos(t) dt for k ∈ {1 ··· , 4n − 1}, be a probability measure on [0, π/2]. Let κ1, κ2 > 0 be such that:

κ1 + κ2 < 1. n Then, there exist U1,U2 ⊂ HP with vol4n(Ui) n = κi, vol4n(HP ) for i = 1, 2 such that:

d(U1,U2) = Sep(µk, κ1, κ2), if and only if we have k = 4m + 3 for m ∈ {0, ··· n − 1}. THE ISOPERIMETRIC INEQUALITY ON COMPACT RANK ONE SYMMETRIC SPACES 35

Proof. If k = 4m + 3 for 1 ≤ m ≤ n − 1, the subset which realizes the separation distance of µk is either a geodesic ball (and its complementary) or a tube around HP k (and its complementary). Indeed, let x ∈ HP n, let r > 0, the ball B(x, r) ⊂ HP n is by definition all the points within distance r to the point x. For k ≥ 1 let HP k be a totally geodesic quaternionic submanifold of dimension k in HP n and for δ > 0, the tube HP k + δ around this submanifold is the set of points within distance δ to HP n. One may consult [10] (for the volume of balls in HP n) and [9] (for the volume of tubes around totally geodesic HP k ⊂ HP n) and verify that: Z r sin(t)4n−1 cos(t)3dt vol (B(x, r)) 4n = 0 . n π/2 vol4n(HP ) Z sin(t)4n−1 cos(t)3dt 0 And

Z r sin(t)4n−4k−1 cos(t)4k+3dt vol ( P k + δ) 4n H = 0 . n π/2 vol4n(HP ) Z sin(t)4n−4k−1 cos(t)4k+3dt 0 n Now suppose k 6= 4m + 3 and there exist U1,U2 ⊂ HP such that for κ1, κ2 > 0 with κ1 + κ2 < 1, we have

vol4n(Ui) n = κi, vol4n(HP ) where i = 1, 2. Moreover, we have:

d(U1,U2) = Sep(µk, κ1, κ2). Let n σ : [0, π/2] → HP , be a geodesic of (maximal) length π/2 parametrized by its arc length t. Let {J1(t), ··· ,J4n(t)} be an orthogonal family of Jacobi fields along σ such that J4n(t) coincides with the tangent vector along σ. We assume for every t ∈ [0, π/2] we have:

q 4n det(< Ji(t),Jk(t) >)i,k=1,··· ,4n = Πi=1kJi(t)k 4n−k+2 k = Cm sin(t) cos(t) .

Let S = {Jj1(t), ··· ,Jjl(t)} be the maximal subset of {J1(t), ··· ,J4n−1(t)} such that Jjβ(0) 6= 0 for every β ∈ {1, ··· , l}. Because of the choice of Jacobi fields, we have l ≥ 1. Indeed if for every i ∈ {1, ··· , 4n − 1}, we have Ji(0) = 0, then n U1 represents a geodesic ball in HP and the needle candidate which realizes the separation distance of a geodesic ball in HP n (and its complementary) is µ = C sin(t)4n−1 cos(t)3dt on a maximal geodesic parametrized by (0, π/2). By definition of Jacobi fields, for every Jjβ ∈ S there exists a (geodesic) variation: n α : [0, π/2] × (−ε, +ε) → HP , such that α is smooth and α(t, 0) = γ(t), 36 YASHAR MEMARIAN and for every s ∈ (−ε, +ε) the curve αs(t) = α(s, t) is a geodesic and such that: ∂α (0, t) = J (t). ∂t jβ

Moreover, by (covariant) differentiating the Jacobi fields Jjβ we have

0 Jjβ(0) = 0. Hence, the image of the variation for time s = 0 lies in a totally geodesic submanifold n of HP . Therefore, the subset U1 is a tube around a totally geodesic submanifold of HP n, i.e. we have a l-dimensional totally geodesic submanifold of HP n such that U1 = N + δ for some δ > 0 and such that: Z δ sin(t)m cos(t)sdt vol (N + δ) 4n = 0 . n π/2 vol4n(HP ) Z sin(t)m cos(t)sdt 0

Since every Jjβ ∈ S is such that the first zero is obtained at distance π/2, i.e. Jjβ(π/2) = 0 and for every t < π/2, we have Jjβ(t) 6= 0, therefore, the totally geodesic submanifold N must be either a point or a HP k ⊂ CP n. Indeed, for totally geodesic submanifolds of HP n which consist of real projective spaces RP l or complex projective space CP l, the first zero of such Jacobi fields are obtained at distance < π/2. The proof hence is completed. 

7.1. End Proof of Theorem 4.3. Since the maximal focal distance of any sub- manifold in HP n is equal to π/2, by applying Lemmas 3.2 and 7.1, for the purpose of estimating the needle separation distance, we can consider only those needle candidates ν such that the density f of their pull-back is defined on the interval (0,L) where L ≤ π/2 and f is of the form presented in proposition 6.1. Hence, let µ1 = f(t)dt be a probability measure on an interval (0,L) where L ≤ π/2 where:

f(t) = C1f1(t)f2(t)f3(t), where

f1(t) = cos(t + a1) ··· cos(t + ap)

f2(t) = cos(t + b1) ··· cos(t + bq)

f3(t) = sin(2t + c1) ··· sin(2t + cs). where C1 > 0, s ≥ 3 and p + q + s = 4n − 1.

For i ∈ {1, ··· max{p, q, s}} we have ai, ci ≥ 0 and bi < 0. Moreover we assume f(0) = f(L) = 0 which is a condition which is verified according to Lemma 7.1. Then, we chose m1, m2, m3 such that

m1 + m2 + m3 = 4n − 1, THE ISOPERIMETRIC INEQUALITY ON COMPACT RANK ONE SYMMETRIC SPACES 37 such that

m2 ≤ p

q ≤ m1

s ≤ m3, and

m1 m2 m3 g(t) = C2 sin(t) cos(t) sin(2t) 4n−4k−1 4k+3 = C2 sin(t) cos(t) , for some k ∈ {0 ··· , n − 1} and some C2 > 0 such that µ2 = g(t)dt is a probability measure on (0, π/2). By applying proposition 6.1 we deduce that for κ1, κ2 > 0 such that

κ1 + κ2 < 1, there exists a k ∈ {1 ··· , n − 1} such that: 4n−4k−1 4k+3 n NHP (κ1, κ2) = Sep(Ck sin(t) cos(t) dt, κ1, κ2). Therefore, according to Proposition 2.2, it remains to find those subsets of HP n which realize the above (optimal) needle separation distance. This is presented in Proposition 7.2. Therefore, we apply the result of Proposition 2.2 and deduce that for every open set U ⊂ HP n, we have:

vol4n(U + δ) ≥ vol4n(B + δ), where B is either a (geodesic) ball or a tube around a totally geodesic HP k ⊂ HP n such that

vol4n(U) = vol4n(B). This ends the proof of Theorem 4.3 for the quaternionic projective spaces.

8. Remarks and Questions Here, we list a few remarks as well as some open problems related to the topic of this paper. • The first question concerns the sharp isoperimetric inequality on the Cayley projective plane CaP 2. We were not able to properly use the geometry of this space in order to study the needles defined upon CaP 2. The main reason for this is the fact that there is not a Riemannian submersion from some round sphere to Cap2, and this makes the study of needle candidates more delicate compared to the cases for RP n, CP n and HP n. It is certainly a very interesting problem of studying needles and the isoperimetry of CaP 2 with the methods of the present paper. • Keeping the spirit of positive curvature and CD(n − 1, n)-needles, it is interesting to study the sharp isoperimetric inequality on Lens spaces. One is invited to study needle candidates on such spaces further and to examine the optimal needle separation distance for such needles. Consult [25] for the study of isoperimetric inequalities on Lens spaces. • Does Klartag’s needle decomposition also hold for non-Riemannian met- ric spaces? This is in comparison with the recent work of Cavalleti and Mondino in [5]. 38 YASHAR MEMARIAN

• Is it possible to generalize Klartag’s partition theorem 2.2 to partitions where the measures (generalized needles) are supported on higher (co)- dimensional sets? A generalization of needles in higher dimension is re- quired and for the case of the round sphere this is provided in [12] and further explained in [20]. • For us, the most interesting problem is to study isoperimetric inequali- ties on (compact) quotients of the Hyperbolic space Hn. The needles on the negatively- curved Riemannian manifolds should (probably) behave as sinhn-concave probability measures. As far as we know, the needles are not yet studied on negatively curved Riemannian manifolds. The needle decomposition tool may be an interesting tool in order to study the famous Cartan-Hadamard conjecture/theorem (as proof of this long standing con- jecture is presented in [8]). This problem consists of isoperimetric inequality on Cartan-Hadamard manifolds.

References [1] D.Bakry, M.Emery,´ Diffusions Hypercontractives. S´eminairede probabilit´esXIX, 1983/84, Lecture Notes in Math, volume 1123, Springer, 177–206, 1985. [2] J.L. Barbosa, M. Do Carmo, J.Eschenburg, Stability of Hypersurfaces of Constant Mean Curvature in Riemannian Manifolds, Mathematische Zeitschrift, volume 197, 123–138, 1988. [3] W.Bench, C. de Franchis, L.Deproit, A.Figalli, S.Gilles, B.Oh, A.Tenne, K.Webster Autour des in´egalit´esisop´erim`etriques,Edited and with a preface by Figalli, Editions de l’Ecole Polytechnique, Palaiseau, 2011. [4] M.Berger, A Panoramic view of Riemannian Geometry, Springer, Berlin New York, Heidel- berg, 2003. [5] F.Cavalletti and A.Mondino, Sharp and Rigid Isoperimetric Inequalities in Metric-Measure Spaces With Lower Ricci Curvature Bounds, Inventiones mathematicae, volume 208, 803–849, 2017. [6] I.Chavel, Isoperimetric Inequalities Differential Geometric and Analytic Perspectives, Cam- bridge Tracts In Mathematics, volume 145, Cambridge University Press, 2001. [7] S.Gallot, D.Hulin and J.Lafontaine, Riemannian Geometry, Universitext, Springer-Verlag, Berlin, Heidelberg, New York, Third Edition, 2012 [8] M.Ghomi and J.Spruck, Total curvature and the isoperimetric inequality in Cartan-Hadamard manifolds, arXiv:1908.09814. [9] X.Gual and A.M. Naveira, The volume of geodesic balls and tubes about totally geodesic sub- manifolds in compact symmetric spaces. Differential Geometry and its Applications, volume 7, 101–113, 1997. [10] A.Gray, Tubes. Progress in Mathematics, Birkhauser, second edition, volume 221, 2001. [11] M.Gromov and V.Milman, Generalization of the spherical isoperimetric inequality to uni- formly convex Banach spaces. compositio.mathematica, tome 62, number 3, 263–282, 1987. [12] M.Gromov, Isoperimetry of Waists and Concentration of Maps. G.A.F.A, Geom. funct. anal. vol 13, 178–215, 2013. [13] M.Gromov, Metric Structures for Riemannian and Non-Riemannian Spaces. Progress in Mathematics, Birkhauser, 2001. [14] M.Gromov, Paul Levy’s Isoperimetric Inequality, Appendex C in Metric Structures for Rie- mannian and Non-Riemannian spaces. [15] M.Gromov, in V.Milman, G.Schechtman, Asymptotic theory of finite dimensional normed spaces, Lecture Notes in Mathematics 1200, Springer-Verlag, Berlin, 1986. [16] L.Guijarro and F.Wilhelm, Focal radius, Rigidity, and Lower Curvature Bounds, Proceedings of the London Mathematical Society, 116,6 , 1519–1552, 2018. [17] R.Hermann, Focal Points of Closed Submanifolds of Riemannian Spaces, Ned. Akad. Wet, 25, 613–628, 1963. [18] B.Klartag, Needle decompositions in Riemannian Geometry, Mem.Amer.Math.Soc., volume 249, number 1180, 2017. THE ISOPERIMETRIC INEQUALITY ON COMPACT RANK ONE SYMMETRIC SPACES 39

[19] B.Klartag, Needle Decompositions and Ricci Curvature, CMC conference: Analysis, Geome- try, and Optimal Transport, 2016. [20] Y.Memarian, On Gromov’s waist of the Sphere Theorem. Journal of Topology and Analysis, volume 3, number 1, 7–36, March 2011. [21] R.Osserman, The Isoperimetric Inequality, Bull. Amer. Math. Soc. 84, 1182–1238, 1978. [22] M.Ritor´eand A.Ros, Stable Constant Mean Curvature Tori and the Isoperimetric Problem in Three Space Forms, Comment. Math. Helvet, 67, 293–305, 1992. [23] A.Ros, The Isoperimetric Problem, Lecture series at the Clay Mathematics Institute Summer School on the Global Theory of Minimal surfaces, summer 2001. [24] T.Sakai, Riemannian Geometry, Translations of Mathematical Monographs, American Math- ematical Society, volume 149, 1996. [25] C.Viana, The isoperimetric problem for lens spaces, Mathematische Annalen, volume 374, 475–497, 2019. [26] C.Villani, Optimal transport, vol. 338 of Grundlehren der Mathematischen Wissenschaften. Springer-Verlag, Berlin, 2009.

Last institution: Department of Mathematics, University of Notre-Dame Email address: [email protected]